You are on page 1of 5

www.advopticalmat.

de
www.MaterialsViews.com
COMMUNICATION

Enabling New Classes of Templated Materials through


Mesoporous Carbon Colloidal Crystals
Matthew D. Goodman, Kevin A. Arpin, Agustin Mihi, Narihito Tatsuda,
Kazuhisa Yano,* and Paul V. Braun*

Colloidal crystals have attracted considerable attention due to colloidal crystals from mesoporous carbon colloidal crystals
their deterministic three-dimensional (3D) structures, inter- which meet these requirements by tailoring the surface charge
esting optical properties and ease of assembly.[1–5] The use of on the mesoporous carbon colloids. We then demonstrate the
colloidal crystals as templates to impart periodic patterns into use of these colloidal crystals as high-surface area templates for
various materials has been broadly employed to create, for traditionally difficult to template materials, employing carbon
example, unique optoelectronic devices,[6,7] sensors,[8–12] and removal processes that are not destructive to the deposited mate-
energy storage devices.[13,14] The general motivation for tem- rials, creating unique, nanostructured inverse opal structures.
plating is to utilize the opals’ interconnected 3D structure to Porous carbon is broadly utilized both for fundamental
define the 3D structure of a material which is inherently difficult studies and large scale commercial applications, including water
to form into a highly regular 3D structure on its own. A single purification, ion exchange,[16] catalysis,[17] conventional battery
replication yields a structure which is an inverse of the colloidal electrodes,[18] emerging battery electrode designs,[19] capacitor
template, and a double replication yields the original structure electrodes,[13] and as a polymer filler. The incorporation of
of the template. This process is only successful if the colloidal small amounts of carbon-black into polymer-based opals have
template can withstand the deposition conditions of the mate- resulted in brilliant colors by absorbing the scattered light.[1]
rial to be templated and there exist conditions whereby the orig- While carbon spheres[20–25] and inverse opals[13,26,27] have been
inal template can be removed without damaging the templated fabricated, carbon opals have only been realized through chem-
material. Given that the most popular template, silica, can only ical vapor deposition (CVD) on a sacrificial, mesoporous silica
be removed with hydrofluoric acid or strong base, chemicals opal.[28] While this creates an opal structure, it would be much
that dissolve many materials, this can be challenging. Polymer more attractive to utilize the self-assembly of carbon colloids,
templates (e.g. polystyrene or poly(methyl methacrylate)) are which would eliminate the CVD and etching steps as well as
easy to remove, but cannot withstand high temperature depo- enable large crack-free structures. Due to carbon's high thermal
sition strategies, limiting their use.[2] Thermally-stable colloids stability (>1000 °C in inert environment),[29] a self-assembled
which could be removed under orthogonal conditions, i.e., con- carbon opal would be an ideal template for materials which can
ditions that do not damage the templated material, would allow only be grown at high-temperature. Additionally, carbon can be
currently inaccessible materials templating strategies. Addition- removed by a simple oxidation step, eliminating the HF etching
ally, if the templates contained additional desirable structural step required to remove silica. Self-assembly of high-quality col-
complexities (e.g. a high surface area) which are replicated in loidal crystals require colloids that are both monodisperse and
the templated material, additional applications may emerge; form a stable suspension, typically accomplished by imparting
for example, dye sensitized solar cells require high-surface area them with a repulsive surface charge[2,30]. By using mesoporous
electrodes,[15] as do many other catalytic devices. In this com- carbon spheres, we can successfully fabricate a high-quality col-
munication, we first demonstrate the fabrication of high-quality loidal crystal. This mesoporous carbon colloidal crystal is then
used as a unique template, due to its high temperature stability,
nanostructured and high surface area, and easy removal. This
M. D. Goodman, K. A. Arpin, Dr. A. Mihi, Prof. P. V. Braun
allows for the fabrication and preservation of unique, nano-
Department of Materials Science and Engineering
Frederick Seitz Materials Research Laboratory structured materials that are inherently difficult to template
Beckman Institute with conventional techniques.
University of Illinois at Urbana-Champaign As a first attempt to make a carbon opal, monodispersed
Urbana, IL, 61801, USA starburst carbon spheres (MSCS), synthesized as previously
E-mail: pbraun@illinois.edu
reported,[31] were dispersed into ethanol and deposited on a
Dr. N. Tatsuda, Dr. K. Yano
Inorganic Materials Laboratory substrate via convective deposition.[7,32,33] Even though the poly-
Toyota Central R & D Labs. Inc dispersity of the MSCS was very low (only 1.038), the result was
Nagakute, Aichi, 480–1192, Japan a disordered film, probably because the zeta-potential of the as
E-mail: k-yano@mosk.tytlabs.co.jp synthesized MSCS was only −14 mV. This compares to a zeta-
Dr. N. Tatsuda, Dr. K. Yano potential of −31 mV measured for typical opal-forming silica par-
Materials Research Laboratory
Toyota Research Institute North America ticles. It has been shown that partial oxidation of carbon fibers
1555 Woodridge Ave. Ann Arbor, MI, 48105, USA introduces ionizable oxygen species (e.g., carboxylic acid);[34]
these species would increase the surface charge. Heating the
DOI: 10.1002/adom.201300120 MSCS in air at 300 and 400 °C increased the surface charge

300 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Optical Mater. 2013, 1, 300–304
www.advopticalmat.de
www.MaterialsViews.com

COMMUNICATION
Table 1. MSCS surface properties after heat treatment for 30 min. the pore volume, size, and specific surface
area are included in Table 1. The structural
Heat treatment Diameter Zeta Potential Pore volume Pore size Specific surface characteristics of the MSCS oxidized at 300
[°C] [nm] [mV] [mL g−1] [nm] area and 400 °C change very little. There is only a
[m2 g−1]
minor decrease in pore volume and a slight
As prepared 484 ± 19 −14 ± 6 1.00 1.67 1670 increase in pore size due to partial collapse of
300 475 ± 7 −26 ± 6 0.92 1.74 1490 micropores, with the surface area remaining
over 1300 m2 g−1. Oxidation at 500 °C resulted
400 473 ± 7 −46 ± 4 0.92 1.87 1360
in significant decreases in both pore volume
500 244 ± 10 −34 ± 4 0.37 - 420 and surface area, and there is now no clear
600 - - - - - mesopore size distribution. To investigate the
surface chemical effects of oxidation, X-ray
with no significant size change; see Table 1 and Figure S1 for Photoelectron Spectroscopy (XPS) was conducted on the as-syn-
zeta-potential and SEM analysis. Significant size reduction thesized and 400 °C treated samples (Table S1 and Figure S2).
occurred at 500 °C, and by 600 °C, the MSCS sample disap- The surface of the oxidized carbon contained significantly more
peared due to complete oxidation of the carbon.[34] To verify the oxygen containing moieties, i.e., hydroxyls, quinones, and car-
porous structure of the MSCS remained through the oxidations, boxylic acids (C-OH, C = O, C-OOH). The presence of these
nitrogen adsorption measurements were conducted (Figure 1); functional groups agree well with the increased surface charge
measured for the oxidized MSCS and is consistent with other
results.[34] Opal formation was then attempted using the MSCS
a with the increased surface charge. Both the 300 °C and 400 °C
700 oxidized samples formed colloidal crystals, with the sample oxi-
As Synthesized
300 °C dized at 400 °C producing the highest quality (shown in Figure 2).
600 400 °C The higher quality of the opals fabricated using the 400 °C oxi-
Volume Adsorbed [cm3 STP g-1]

500 °C
dized MSCS can be attributed to the increased stability of the
500
suspension due to the higher surface charge compared to the as-
synthesized and 300 °C oxidized (−46 mV vs. −14 and −26 mV,
400
respectively). Although opals produced this way show high
degree of order in the Scanning Electron Microscopy (SEM)
300
micrographs, due to the strong absorption of carbon, a low
optical reflectivity and transmission was observed (Figure S3).
200
Due to the MSCS porosity and potential for complete
removal by oxidation in air, the carbon colloidal crystals are
100
ideally suited as templates to create unique nanostructures
after inversion. Employed here are two gas-phase deposition
0
0.0 0.2 0.4 0.6 0.8 1.0 techniques that allow deep infilling of the mesopores: atomic
Relative Pressure, P/P0 layer deposition (ALD) of hafnia (HfO2) and alumina (Al2O3)
and static chemical vapor deposition (CVD) of silicon (Si). SEM
b 1.0
As synthesized
0.9 300 °C
0.8 400 °C
500 °C
dV(w) [cm3 nm-1 g-1]

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 1 2 3 4 5
Pore Width [nm]

Figure 1. (a) Nitrogen adsorption (solid circles) and desorption (open


circles) isotherms of as prepared MSCS (black curve), and MSCS oxidized Figure 2. SEM micrographs showing cross-section (a,b) and oblique
at 300 °C (red), 400 °C (blue), and 500 °C (green). (b) Calculated pore (c) views of 5 layer opals, and (d) cross-section micrograph of 7 layer
size distribution from (a). opal, all fabricated from 400 °C oxidized MSCS.

Adv. Optical Mater. 2013, 1, 300–304 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 301
www.advopticalmat.de
www.MaterialsViews.com
COMMUNICATION

To verify these hypotheses, experiments


were conducted to crystallize the HfO2 prior
to MSCS template removal. As deposited,
the HfO2 is amorphous. To crystallize the
HfO2 prior to carbon removal, the sample
was annealed in forming gas (5% H2 in Ar)
at 600 °C for 1 h. This replicates the thermal
removal process time and temperature; how-
ever, the forming gas limits MSCS oxidation.
Crystallization of the HfO2 prior to template
removal is only possible due to carbon’s
high thermal stability and inert nature with
respect to HfO2. After annealing, the MSCS
was removed through oxygen plasma. Cross-
sectional SEM (Figure 3d) of this sample is
strikingly similar to the as-deposited, oxygen
plasma removed samples (Figure 3c). It
appears the MSCS provided a support and
template during HfO2 crystallization, lim-
Figure 3. SEM and TEM (insets) micrographs of HfO2-MSCS system. (a) FIB cross-section of iting sintering and grain growth, preserving
a carbon opal with HfO2 infiltrating mesopores (white inset: contrast enhanced). (b-d) Fracture the templated nanostructure. X-ray diffrac-
surfaces (b) HfO2 inverse opal created after carbon opal removal at 600 °C in air. (c) HfO2 tion (XRD) on the as-deposited and annealed
inverse opal fabricated through carbon removal via oxygen plasma. (d) HfO2 inverse opal samples was used to confirm the amorphous
annealed at 600 °C with 5% H2 in Ar, with the carbon removed via oxygen plasma. nature of the as deposed HfO2 and its crys-
talline nature after annealing. For this experi-
micrographs after 100 cycles of ALD HfO2, 10 nm nominally ment, it is important that the majority of the HfO2 is depos-
on a flat substrate, are shown in Figure 3. Verification of pore ited in the pores of the MSCS and not on the MSCS surface
infiltration was done through a Focused Ion Beam (FIB) cut (the x-ray experiment does not distinguish between HfO2 on
on a carbon/HfO2 composite opal, i.e., prior to carbon removal the surface of the MSCS surface and in the pores), and so the
(Figure 3a). Transmission Electron Microscopy (TEM) was also number of ALD cycles was reduced to 60, depositing nominally
conducted on the composite colloid (Figure 3a, top). The top 6 nm of HfO2. Most of the HfO2 will be inside the MSCS, and
of the opal has a thick sputtered gold coating required for FIB only a few nm will be on the surface of the carbon colloids.
milling; no gold was coated after milling on the exposed sur- XRD shows the as-deposited HfO2 is amorphous, while the
face. Carbon, due to its low atomic number, appears darker annealed sample is crystalline (Figure S4). From the Scherrer
than the HfO2 in the SEM micrograph. The micrograph shows Equation (Equation S1), the average crystallite was found to be
a bright HfO2 shell and spokes of HfO2 penetrating the MSCS. 8.4 nm, substantially greater than the 1.87 nm MSCS pore size
Measurements reveal the HfO2 infiltrates approximately but significantly less than the total length infiltrated into the
80 nm. ALD is known to fill deep vias and other high aspect pores (∼80 nm), indicating the HfO2 crystallites may be rod-
ratio structures, and thus deep infilling of the MSCS was not like. Transmission electron microscopy (TEM) (Figure 3d, top)
a surprise.[35] and selected area electron diffraction (SAED, Figure S5) con-
After a brief Reactive Ion Etch (RIE) to open the top HfO2 sur- firmed the crystalline nature of HfO2.
face, carbon was thermally removed at 600 °C for 1 h. Unlike 100 cycles (10 nm nominally) of ALD Al2O3, with the MSCS
wet or deep RIE etching required for silica template removal, thermally removed, is shown with a FIB cut in Figure 4a. As
this thermal removal process does not remove HfO2, so all the with the HfO2, the thermal removal process is completely
deposited HfO2 remains. Figure 3b shows the HfO2 inverse opal orthogonal to Al2O3 removal. In most of the MSCS, the Al2O3
SEM and TEM micrographs. In these, a granular HfO2 structure infiltrates into the center; however, 'defects' of partially unfilled
exists, perhaps due to the HfO2 crystallizing during the thermal MSCS occur (top right MSCS in Figure 4a). The TEM micro-
MSCS removal. Since carbon can be readily removed through graph of the Al2O3 inverse colloid in Figure 4b shows the solid
oxygen plasma at room temperature, it was possible to evaluate Al2O3 shell and porous center. Unlike for HfO2, the Al2O3
the structure of the HfO2 in the as-deposited state. The top of the inverse opal did not crystallize, and its mesostructure did not
HfO2-coated MSCS opal was opened via RIE etching, followed appear to Oswald ripen during thermal carbon removal at 600 °C
by an oxygen plasma etch to remove the MSCS. The resulting (XRD in Figure S4). This may not be surprising, as 1100 °C is a
inverse opal (Figure 3c) contrasts sharply with the thermally typical crystallization temperature for alumina.[36]
removed MSCS; a smooth, instead of granular, nanostructure is The final successful material investigated for template inver-
observed. It appears that during the MSCS thermal removal pro- sion was Si, which as grown using static CVD. SEM and TEM
cess, grain growth and sintering in the HfO2 occur simultane- micrographs are shown in Figure 4c,d. To prevent Si oxidation,
ously with carbon removal. Once the MSCS support is removed, MSCS were removed via room temperature oxygen plasma
the HfO2 is free to crystallize and coarsen, while the as-deposited after removing the top Si over layer via RIE. The conformal Si
HfO2 better replicates the ultra-high surface area MSCS. static CVD deposition infiltrates the MSCS template, creating

302 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Optical Mater. 2013, 1, 300–304
www.advopticalmat.de
www.MaterialsViews.com

COMMUNICATION
at a wavelength of 840 nm. A MSCS opal was also coated with
300 cycles HfO2 (30 nm nominally). The optical measurements
are shown in Figure 5b after the MSCS were removed. 30 nm of
HfO2 is below the pinch-off point (when the colloid interstitials
fill and block further precursor deposition) of 37 nm for the
479 nm colloids. For the thicker HfO2 deposition, the reflection
peak red-shifts to 1.0 μm and increases to 35% with well-defined
Fabry-Perot fringes. Since the mesopores are already full after
100 ALD cycles, the extra 200 cycles simply increase the thick-
ness of solid HfO2 shell around each MSCS particle, leading
to the red-shift and increases in intensity of the reflection
peak. The Al2O3 optical measurements are shown in Figure 5c.
Interestingly, no reflection peak is observed; this could be
the result of the infilling 'defects', where a portion of the col-
loids have large voids in the center. These defects perhaps act
Figure 4. (a) FIB cross-section cut of ALD Al2O3 inverse opal with carbon as strong scattering centers. Optical measurements of the Si
template thermally removed. (b) TEM micrograph of Al2O3 structure in inverse structure are shown in Figure 5d; the main reflectance
(a). (c) Silicon inverse opal fabricated via static CVD; carbon opal tem- peak is at 940 nm with 19% reflectance. The transmittance
plate was removed via oxygen plasma. (d) TEM micrograph of silicon
measurement shows a primary dip of 63% corresponding with
structures in (c).
the reflectance peak; the remaining 18% of the light is either
absorbed or scattered. Interestingly, the Si inverse structure has
a porous-Si interior. As expected, XRD (Figure S4) reveals the a transmittance of near 95% at longer wavelengths, compared
as-grown Si is amorphous, as Si crystallization requires heating to only ∼55% for a typical Si inverse opal.[37] The high transmit-
to 1000 °C for several hours. Platinum ALD in the MSCS was tance (low reflectance) is evidence that the effective refractive
also attempted. However, due to oxygen being a reactant in the index of the porous Si structure is much lower than pure Si
ALD growth, the MSCS opal oxidizes and loses its well-defined (∼3.5).
order. In conclusion, we have shown that by tailoring the surface
Optical measurements were conducted on HfO2, Al2O3, and charge of carbon colloids, high-quality carbon colloidal crys-
Si inverted opals (Figure 5). Figure 5a shows the 100 cycle ALD tals can be fabricated and that these high-surface area colloidal
HfO2 inverse structure with a primary reflection peak of 16% crystals can be used as templates for a variety of materials,

a 1.0 1.0 b 1.0 1.0


0.9 0.9 0.9 0.9
0.8 0.8 0.8 0.8
0.7 0.7 0.7 0.7
Transmittance
Reflectance

Reflectance

0.6 0.6 0.6 0.6 Transmittance


0.5 0.5 0.5 0.5
0.4 0.4 0.4 0.4
0.3 0.3 0.3 0.3
0.2 0.2 0.2 0.2
0.1 0.1 0.1 0.1
0.0 0.0 0.0 0.0
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6

Wavelength (μ m) Wavelength (μm)

c 1.0 1.0 d 1.0 1.0


0.9 0.9 0.9 0.9
0.8 0.8 0.8 0.8
0.7 0.7 0.7 0.7
Transmittance

Transmittance
Reflectance

Reflectance

0.6 0.6 0.6 0.6


0.5 0.5 0.5 0.5
0.4 0.4 0.4 0.4
0.3 0.3 0.3 0.3
0.2 0.2 0.2 0.2
0.1 0.1 0.1 0.1
0.0 0.0 0.0 0.0
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6

Wavelength (μm) Wavelength (μm)

Figure 5. Optical measurements on inverse opals structures after carbon removal: (a) 100 cycle ALD HfO2; (b) 300 cycle ALD HfO2; (c) 100 cycle ALD
Al2O3; (d) static CVD silicon.

Adv. Optical Mater. 2013, 1, 300–304 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 303
www.advopticalmat.de
www.MaterialsViews.com
COMMUNICATION

creating unique, high-surface area nanostructured inverse [4] C. I. Aguirre, E. Reguera, A. Stein, Adv. Funct. Mater. 2011, 21,
opals. These materials, HfO2, Al2O3, and Si, penetrate deep into 210.
the mesopores of the MSCS. Importantly, the carbon removal [5] S.-H. Kim, S. Y. Lee, S.-M. Yang, G.-R. Yi, Asia Mater. 2011, 3, 25.
process is completely orthogonal to the deposited material [6] E. C. Nelson, N. L. Dias, K. P. Bassett, S. N. Dunham, V. Verma,
M. Miyake, P. Wiltzius, J. A. Rogers, J. J. Coleman, X. Li, P. V. Braun,
removal, preserving the fine-scale mesostructure of the MSCS
Nat. Mater. 2011, 10, 676.
in the templated materials. Because MSCS removal is orthog- [7] A. Mihi, C. Zhang, P. V. Braun, Angew. Chem. Int. Ed. 2011, 50,
onal to the etching of many materials, MSCS provide a general 5712.
approach for templating materials which currently cannot be [8] N. Griffete, H. Frederich, A. Maitre, S. Ravaine, M. M. Chehimi,
templated by conventional opals. C. Mangeney, Langmuir 2012, 28, 1005.
[9] Z. Pan, J. Ma, J. Yan, M. Zhou, J. Gao, J. Mater. Chem. 2012, 22,
2018.
Experimental Section [10] R. A. Barry, P. Wiltzius, Langmuir 2006, 22, 1369.
[11] Y. J. Lee, P. V. Braun, Adv. Mater. 2003, 15, 563.
Synthesis: Monodisperse Starburst Carbon Spheres (MSCS) were
[12] J. Wang, Y. Cao, Y. Feng, F. Yin, J. Gao, Adv. Mater. 2007, 19,
synthesized as previously reported.[31] The MSCS were oxidized in a
3865.
Lindberg Furnace using a 30 min ramp, 30 min hold, and cooled to room
temperature. For opal fabrication, Piranha-cleaned substrates (glass or [13] Y. Isshiki, M. Nakamura, S. Tabata, K. Dokko, M. Watanabe, Polym.
quartz) were placed at a 20° angle in a 20 mL scintillation vial with 0.7 g Advan. Technol. 2011, 22, 1254.
colloidal suspension (0.5–2 wt% in ethanol). The vials were placed in an [14] H. Zhang, X. Yu, P. V. Braun, Nat. Nanotechnol. 2011, 6, 277.
incubator (Fisher, Isotemp 125D) and held at 40 °C overnight. [15] B. Mandlmeier, J. M. Szeifert, D. Fattakhova-Rohlfing, H. Amenitsch,
Infilltration and Template Removal: HfO2 and Al2O3 ALD, using T. Bein, J. Am. Chem. Soc. 2011, 133, 17274.
a Cambridge Nanotech ALD system, were done at 200 °C and 80 °C, [16] K. R. Benak, L. Dominguez, J. Economy, C. L. Mangun, Carbon 2002,
respectively, on fabricated carbon opals with both recipes having a 40, 2323.
growth rate of 1 Å cycle−1. Si static CVD was done at 350 °C for 2 h using [17] J. Zhang, Y. Zhang, S. Lian, Y. Liu, Z. Kang, S.-T. Lee, J. Colloid Interf.
Si2H6.[38] A Reactive Ion Etch (RIE) using O2 and CF4 gasses (1 sccm Sci. 2011, 361, 503.
each, 10 mTorr, 75 W, 1 nm min−1 HfO2 removal) was done to expose [18] F. Cheng, Z. Tao, J. Liang, J. Chen, Chem. Mater. 2008, 20,
the MSCS for thermal or oxygen plasma removal. Thermal removal was 667.
done at 600 °C for 1 h with a 30 min ramp. Oxygen plasma removal was [19] J. Xiao, D. Mei, X. Li, W. Xu, D. Wang, G. L. Graff, W. D. Bennett,
done using 20 sccm O2 at 400 mTorr, 200 W, for 2 h. Z. Nie, L. V. Saraf, I. A. Aksay, J. Liu, J. G. Zhang, Nano Lett. 2011,
Characterization: Nitrogen physisorption isotherms, BET specific 11, 5071.
surface area, and pore volume were measured and calculated as [20] J. Liu, S. Z. Qiao, H. Liu, J. Chen, A. Orpe, D. Zhao, G. Q. Lu, Angew.
previously described.[31] Scanning electron microscopy (SEM) was Chem.-Int. Ed. 2011, 50, 5947.
done on Hitachi S-4700 or S-4800. Transmission electron microscopy
[21] X. M. Sun, Y. D. Li, Angew. Chem.-Int. Ed. 2004, 43, 597.
(TEM) was done on a JEOL 2010LAB6 at an accelerating voltage of
[22] S. B. Yoon, K. Sohn, J. Y. Kim, C. H. Shin, J. S. Yu, T. Hyeon, Adv.
200 kV. Focus ion beam (FIB) milling was done on a FEI Beam 235 FIB.
Mater. 2002, 14, 19.
Zeta-potential measurements were conducted in Millipore water on a
NICOMP 380 ZLS Particle Sizer. [23] M. B. Zheng, J. M. Cao, X. Chang, J. Wang, J. S. Liu, X. J. Ma, Mater.
Lett. 2006, 60, 2991.
[24] J. P. Paraknowitsch, A. Thomas, M. Antonietti, Chem. Mater. 2009,
Supporting Information 21, 1170.
[25] C. Y. Chen, X. D. Sun, X. C. Jiang, D. Niu, A. B. Yu, Z. G. Liu, J. G. Li,
Supporting Information is available from the Wiley Online Library or Nanoscale Res. Lett. 2009, 4, 971.
from the author.
[26] A. A. Zakhidov, R. H. Baughman, Z. Iqbal, C. X. Cui, I. Khayrullin,
S. O. Dantas, I. Marti, V. G. Ralchenko, Science 1998, 282,
Acknowledgements 897.
[27] J. Lee, J. Kim, T. Hyeon, Adv. Mater. 2006, 18, 2073.
This work was supported by Toyota Central R&D Labs (fabrication and [28] Y. Yamada, M. Ishii, T. Nakamura, K. Yano, Langmuir 2010, 26,
structural characterization) and the US Department of Energy ‘Light
10044.
Material Interactions in Energy Conversion’ Energy Frontier Research
[29] A. C. Ferrari, B. Kleinsorge, N. A. Morrison, A. Hart, V. Stolojan,
Center under grant DE-SC0001293 (optical studies). The authors would
J. Robertson, J. Appl. Phys. 1999, 85, 7191.
like to thank Dr. J. Cho, H. Ning, and S. Kranz for technical assistance.
This work was carried out in part in the Frederick Seitz Materials [30] S. Wong, V. Kitaev, G. A. Ozin, J. Am. Chem. Soc. 2003, 125,
Research Laboratory Central Facilities, University of Illinois. 15589.
[31] T. Nakamura, Y. Yamada, K. Yano, Micropor. Mesopor. Mat. 2009,
Received: February 21, 2013 117, 478.
Published online: March 21, 2013 [32] H. T. Yang, P. Jiang, J. Colloid Interf. Sci. 2010, 352, 558.
[33] B. G. Prevo, O. D. Velev, Langmuir 2004, 20, 2099.
[34] C. L. Mangun, K. R. Benak, M. A. Daley, J. Economy, Chem. Mater.
[1] C. E. Finlayson, P. Spahn, D. R. Snoswell, G. Yates, A. Kontogeorgos, 1999, 11, 3476.
A. I. Haines, G. P. Hellmann, J. J. Baumberg, Adv. Mater. 2011, 23, [35] M. Koyanagi, T. Fukushima, T. Tanaka, Proc. IEEE 2009, 97, 49.
1540. [36] P. S. Santos, H. S. Santos, S. P. Toledo, Mater. Res. 2000, 3, 104.
[2] F. Marlow, P. Sharifi, R. Brinkmann, C. Mendive, Angew. Chem. Int. [37] G. von Freymann, V. Kitaev, B. V. Lotsch, G. A. Ozin, Chem. Soc. Rev.
Ed. 2009, 48, 6212. 2013, DOI: 10.1039/C2CS35309A.
[3] H. Yamada, T. Nakamura, Y. Yamada, K. Yano, Adv. Mater. 2009, 21, [38] S. A. Rinne, F. Garcia-Santamaria, P. V. Braun, Nat. Photonics 2008,
4134. 2, 52.

304 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Optical Mater. 2013, 1, 300–304

You might also like